リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

大学・研究所にある論文を検索できる 「Motif-centric phosphoproteomics to target kinase-mediated signaling pathways」の論文概要。リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

コピーが完了しました

URLをコピーしました

論文の公開元へ論文の公開元へ
書き出し

Motif-centric phosphoproteomics to target kinase-mediated signaling pathways

Tsai, Chia-Feng Ogata, Kosuke Sugiyama, Naoyuki Ishihama, Yasushi 京都大学 DOI:10.1016/j.crmeth.2021.100138

2022.01

概要

Identifying cellular phosphorylation pathways based on kinase-substrate relationships is a critical step to understanding the regulation of physiological functions in cells. Mass spectrometry-based phosphoproteomics workflows have made it possible to comprehensively collect information on individual phosphorylation sites in a variety of samples. However, there is still no generic approach to uncover phosphorylation networks based on kinase-substrate relationships in rare cell populations. Here, we describe a motif-centric phosphoproteomics approach combined with multiplexed isobaric labeling, in which in vitro kinase reactions are used to generate targeted phosphopeptides, which are spiked into one of the isobaric channels to increase detectability. Proof-of-concept experiments demonstrate selective and comprehensive quantification of targeted phosphopeptides by using multiple kinases for motif-centric channels. More than 7, 000 tyrosine phosphorylation sites were quantified from several tens of micrograms of starting materials. This approach enables the quantification of multiple phosphorylation pathways under physiological or pathological regulation in a motif-centric manner.

この論文で使われている画像

参考文献

Abe, Y., Nagano, M., Tada, A., Adachi, J., and Tomonaga, T. (2017). Deep

phosphotyrosine proteomics by optimization of phosphotyrosine enrichment

and MS/MS parameters. J. Proteome Res. 16, 1077–1086.

Bekker-Jensen, D.B., Martinez-Val, A., Steigerwald, S., Ruther, P., Fort, K.L.,

Arrey, T.N., Harder, A., Makarov, A., and Olsen, J.V. (2020). A compact quadrupole-orbitrap mass spectrometer with FAIMS interface improves proteome

coverage in short LC gradients. Mol. Cell. Proteomics 19, 716–729.

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

Article

Bian, Y., Li, L., Dong, M., Liu, X., Kaneko, T., Cheng, K., Liu, H., Voss, C., Cao,

X., Wang, Y., et al. (2016). Ultra-deep tyrosine phosphoproteomics enabled by

a phosphotyrosine superbinder. Nat. Chem. Biol. 12, 959–966.

OPEN ACCESS

Hunter, T. (2009). Tyrosine phosphorylation: thirty years and counting. Curr.

Opin. Cell Biol. 21, 140–146.

Bloom, J., and Cross, F.R. (2007). Multiple levels of cyclin specificity in cell-cycle control. Nat. Rev. Mol. Cell Biol. 8, 149–160.

Hunter, T., and Sefton, B.M. (1980). Transforming gene product of Rous sarcoma virus phosphorylates tyrosine. Proc. Natl. Acad. Sci. U S A 77, 1311–

1315.

Budnik, B., Levy, E., Harmange, G., and Slavov, N. (2018). SCoPE-MS: mass

spectrometry of single mammalian cells quantifies proteome heterogeneity

during cell differentiation. Genome Biol. 19, 161.

Imamura, H., Sugiyama, N., Wakabayashi, M., and Ishihama, Y. (2014). Largescale identification of phosphorylation sites for profiling protein kinase selectivity. J. Proteome Res. 13, 3410–3419.

Cheung, T.K., Lee, C.Y., Bayer, F.P., McCoy, A., Kuster, B., and Rose, C.M.

(2021). Defining the carrier proteome limit for single-cell proteomics. Nat.

Methods 18, 76–83.

Imamura, H., Wagih, O., Niinae, T., Sugiyama, N., Beltrao, P., and Ishihama, Y.

(2017). Identifications of putative PKA substrates with quantitative phosphoproteomics and primary-sequence-based scoring. J. Proteome Res. 16,

1825–1830.

Chon, H.J., Bae, K.J., Lee, Y., and Kim, J. (2015). The casein kinase 2 inhibitor,

CX-4945, as an anti-cancer drug in treatment of human hematological malignancies. Front. Pharmacol. 6, 70.

Chua, X.Y., Mensah, T., Aballo, T., Mackintosh, S.G., Edmondson, R.D., and

Salomon, A.R. (2020). Tandem mass tag approach utilizing pervanadate

BOOST channels delivers deeper quantitative characterization of the tyrosine

phosphoproteome. Mol. Cell. Proteomics 19, 730–743.

Iwasaki, M., Miwa, S., Ikegami, T., Tomita, M., Tanaka, N., and Ishihama, Y.

(2010). One-dimensional capillary liquid chromatographic separation coupled

with tandem mass spectrometry unveils the Escherichia coli proteome on a microarray scale. Anal. Chem. 82, 2616–2620.

Cohen, P. (2002). The origins of protein phosphorylation. Nat. Cell Biol. 4,

E127–E130.

Li, X., Cox, J.T., Huang, W., Kane, M., Tang, K., and Bieberich, C.J. (2016).

Quantifying kinase-specific phosphorylation stoichiometry using stable

isotope labeling in a reverse in-gel kinase assay. Anal. Chem. 88, 11468–

11475.

Cox, J., and Mann, M. (2008). MaxQuant enables high peptide identification

rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372.

Masuda, T., Tomita, M., and Ishihama, Y. (2008). Phase transfer surfactantaided trypsin digestion for membrane proteome analysis. J. Proteome Res.

7, 731–740.

Dong, M., Bian, Y., Wang, Y., Dong, J., Yao, Y., Deng, Z., Qin, H., Zou, H., and

Ye, M. (2017). Sensitive, robust, and cost-effective approach for tyrosine

phosphoproteome analysis. Anal. Chem. 89, 9307–9314.

McAlister, G.C., Nusinow, D.P., Jedrychowski, M.P., Wuhr, M., Huttlin, E.L.,

Erickson, B.K., Rad, R., Haas, W., and Gygi, S.P. (2014). MultiNotch MS3 enables accurate, sensitive, and multiplexed detection of differential expression

across cancer cell line proteomes. Anal. Chem. 86, 7150–7158.

Dou, M., Clair, G., Tsai, C.F., Xu, K., Chrisler, W.B., Sontag, R.L., Zhao, R.,

Moore, R.J., Liu, T., Pasa-Tolic, L., et al. (2019). High-throughput single cell

proteomics enabled by multiplex isobaric labeling in a nanodroplet sample

preparation platform. Anal. Chem. 91, 13119–13127.

Erickson, B.K., Mintseris, J., Schweppe, D.K., Navarrete-Perea, J., Erickson,

A.R., Nusinow, D.P., Paulo, J.A., and Gygi, S.P. (2019). Active instrument

engagement combined with a real-time database search for improved performance of sample multiplexing workflows. J. Proteome Res. 18, 1299–1306.

Fang, B., Izumi, V., Rix, L.L.R., Welsh, E., Pike, I., Reuther, G.W., Haura, E.B.,

Rix, U., and Koomen, J.M. (2020). Lowering sample requirements to study

tyrosine kinase signaling using phosphoproteomics with the TMT calibrator

approach. Proteomics 20, e2000116.

Gallien, S., Kim, S.Y., and Domon, B. (2015). Large-scale targeted proteomics

using internal standard triggered-parallel reaction monitoring (IS-PRM). Mol.

Cell. Proteomics 14, 1630–1644.

Hebert, A.S., Prasad, S., Belford, M.W., Bailey, D.J., McAlister, G.C., Abbatiello, S.E., Huguet, R., Wouters, E.R., Dunyach, J.J., Brademan, D.R., et al.

(2018). Comprehensive single-shot proteomics with FAIMS on a hybrid Orbitrap mass spectrometer. Anal. Chem. 90, 9529–9537.

Hogrebe, A., von Stechow, L., Bekker-Jensen, D.B., Weinert, B.T., Kelstrup,

C.D., and Olsen, J.V. (2018). Benchmarking common quantification strategies

for large-scale phosphoproteomics. Nat. Commun. 9, 1045.

Hornbeck, P.V., Zhang, B., Murray, B., Kornhauser, J.M., Latham, V., and

Skrzypek, E. (2015). PhosphoSitePlus, 2014: mutations, PTMs and recalibrations. Nucleic Acids Res. 43, D512–D520.

Huang da, W., Sherman, B.T., and Lempicki, R.A. (2009). Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat.

Protoc. 4, 44–57.

Hughes, C.S., Zhu, C., Spicer, V., Krokhin, O.V., and Morin, G.B. (2017). Evaluating the characteristics of reporter ion signal acquired in the Orbitrap

analyzer for isobaric mass tag proteome quantification experiments.

J. Proteome Res. 16, 1831–1838.

Humphrey, S.J., Azimifar, S.B., and Mann, M. (2015). High-throughput phosphoproteomics reveals in vivo insulin signaling dynamics. Nat. Biotechnol.

33, 990–995.

Hunter, T. (2000). Signaling–2000 and beyond. Cell 100, 113–127.

Mertins, P., Tang, L.C., Krug, K., Clark, D.J., Gritsenko, M.A., Chen, L.,

Clauser, K.R., Clauss, T.R., Shah, P., Gillette, M.A., et al. (2018). Reproducible

workflow for multiplexed deep-scale proteome and phosphoproteome analysis of tumor tissues by liquid chromatography-mass spectrometry. Nat. Protoc. 13, 1632–1661.

Moriya, Y., Kawano, S., Okuda, S., Watanabe, Y., Matsumoto, M., Takami, T.,

Kobayashi, D., Yamanouchi, Y., Araki, N., Yoshizawa, A.C., et al. (2019). The

jPOST environment: an integrated proteomics data repository and database.

Nucleic Acids Res. 47, D1218–D1224.

Mundina-Weilenmann, C., Chang, C.F., Gutierrez, L.M., and Hosey, M.M.

(1991). Demonstration of the phosphorylation of dihydropyridine-sensitive calcium channels in chick skeletal muscle and the resultant activation of the channels after reconstitution. J. Biol. Chem. 266, 4067–4073.

Needham, E.J., Parker, B.L., Burykin, T., James, D.E., and Humphrey, S.J.

(2019). Illuminating the dark phosphoproteome. Sci. Signal. 12, eaau8645.

O’Shea, J.P., Chou, M.F., Quader, S.A., Ryan, J.K., Church, G.M., and

Schwartz, D. (2013). pLogo: a probabilistic approach to visualizing sequence

motifs. Nat. Methods 10, 1211–1212.

Ogata, K., and Ishihama, Y. (2020). Extending the separation space with trapped ion mobility spectrometry improves the accuracy of isobaric tag-based

quantitation in proteomic LC/MS/MS. Anal. Chem. 92, 8037–8040.

Ogata, K., Tsai, C.F., and Ishihama, Y. (2021). Nanoscale solid-phase isobaric

labeling for multiplexed quantitative phosphoproteomics. J. Proteome Res.

20, 4193–4202.

Possemato, A.P., Paulo, J.A., Mulhern, D., Guo, A., Gygi, S.P., and Beausoleil,

S.A. (2017). Multiplexed phosphoproteomic profiling using titanium dioxide

and immunoaffinity enrichments reveals complementary phosphorylation

events. J. Proteome Res. 16, 1506–1514.

Rappsilber, J., Mann, M., and Ishihama, Y. (2007). Protocol for micro-purification, enrichment, pre-fractionation and storage of peptides for proteomics using StageTips. Nat. Protoc. 2, 1896–1906.

Russo, G.L., Vandenberg, M.T., Yu, I.J., Bae, Y.S., Franza, B.R., Jr., and

Marshak, D.R. (1992). Casein kinase II phosphorylates p34cdc2 kinase in G1

phase of the HeLa cell division cycle. J. Biol. Chem. 267, 20317–20325.

Cell Reports Methods 2, 100138, January 24, 2022 11

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Sharma, K., D’Souza, R.C., Tyanova, S., Schaab, C., Wisniewski, J.R., Cox, J.,

and Mann, M. (2014). Ultradeep human phosphoproteome reveals a distinct

regulatory nature of Tyr and Ser/Thr-based signaling. Cell Rep. 8, 1583–1594.

Songyang, Z., Lu, K.P., Kwon, Y.T., Tsai, L.H., Filhol, O., Cochet, C., Brickey,

D.A., Soderling, T.R., Bartleson, C., Graves, D.J., et al. (1996). A structural basis for substrate specificities of protein Ser/Thr kinases: primary sequence

preference of casein kinases I and II, NIMA, phosphorylase kinase, calmodulin-dependent kinase II, CDK5, and Erk1. Mol. Cell. Biol. 16, 6486–6493.

Stokes, M.P., Farnsworth, C.L., Moritz, A., Silva, J.C., Jia, X., Lee, K.A., Guo,

A., Polakiewicz, R.D., and Comb, M.J. (2012). PTMScan direct: identification

and quantification of peptides from critical signaling proteins by immunoaffinity enrichment coupled with LC-MS/MS. Mol. Cell. Proteomics 11, 187–201.

Stokes, M.P., Farnsworth, C.L., Gu, H., Jia, X., Worsfold, C.R., Yang, V., Ren,

J.M., Lee, K.A., and Silva, J.C. (2015). Complementary PTM profiling of drug

response in human gastric carcinoma by immunoaffinity and IMAC methods

with total proteome analysis. Proteomes 3, 160–183.

Article

teomic enrichment through complementary metal-directed immobilized metal

ion affinity chromatography. Anal. Chem. 86, 685–693.

Tsai, C.F., Wang, Y.T., Yen, H.Y., Tsou, C.C., Ku, W.C., Lin, P.Y., Chen, H.Y.,

Nesvizhskii, A.I., Ishihama, Y., and Chen, Y.J. (2015). Large-scale determination of absolute phosphorylation stoichiometries in human cells by motif-targeting quantitative proteomics. Nat. Commun. 6, 6622.

Tsai, C.F., Zhao, R., Williams, S.M., Moore, R.J., Schultz, K., Chrisler, W.B.,

Pasa-Tolic, L., Rodland, K.D., Smith, R.D., Shi, T., et al. (2020). An improved

boosting to amplify signal with isobaric labeling (iBASIL) strategy for precise

quantitative single-cell proteomics. Mol. Cell. Proteomics 19, 828–838.

Tyanova, S., Temu, T., and Cox, J. (2016a). The MaxQuant computational platform for mass spectrometry-based shotgun proteomics. Nat. Protoc. 11,

2301–2319.

Tyanova, S., Temu, T., Sinitcyn, P., Carlson, A., Hein, M.Y., Geiger, T., Mann,

M., and Cox, J. (2016b). The Perseus computational platform for comprehensive analysis of (prote)omics data. Nat. Methods 13, 731–740.

Stopfer, L.E., Conage-Pough, J.E., and White, F.M. (2021a). Quantitative consequences of protein carriers in immunopeptidomics and tyrosine phosphorylation MS(2) analyses. Mol. Cell. Proteomics 20, 100104.

Villen, J., Beausoleil, S.A., Gerber, S.A., and Gygi, S.P. (2007). Large-scale

phosphorylation analysis of mouse liver. Proc. Natl. Acad. Sci. U S A 104,

1488–1493.

Stopfer, L.E., Flower, C.T., Gajadhar, A.S., Patel, B., Gallien, S., Lopez-Ferrer,

D., and White, F.M. (2021b). High-density, targeted monitoring of tyrosine

phosphorylation reveals activated signaling networks in human tumors. Cancer Res. 81, 2495–2509.

Wang, Y.T., Pan, S.H., Tsai, C.F., Kuo, T.C., Hsu, Y.L., Yen, H.Y., Choong,

W.K., Wu, H.Y., Liao, Y.C., Hong, T.M., et al. (2017). Phosphoproteomics reveals HMGA1, a CK2 substrate, as a drug-resistant target in non-small cell

lung cancer. Sci. Rep. 7, 44021.

Sugiyama, N., Imamura, H., and Ishihama, Y. (2019). Large-scale discovery of

substrates of the human kinome. Sci. Rep. 9, 10503.

Xue, L., Wang, W.H., Iliuk, A., Hu, L., Galan, J.A., Yu, S., Hans, M., Geahlen,

R.L., and Tao, W.A. (2012). Sensitive kinase assay linked with phosphoproteomics for identifying direct kinase substrates. Proc. Natl. Acad. Sci. U S A 109,

5615–5620.

Szklarczyk, D., Gable, A.L., Lyon, D., Junge, A., Wyder, S., Huerta-Cepas, J.,

Simonovic, M., Doncheva, N.T., Morris, J.H., Bork, P., et al. (2019). STRING

v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids

Res. 47, D607–D613.

Thompson, A., Schafer, J., Kuhn, K., Kienle, S., Schwarz, J., Schmidt, G., Neumann, T., Johnstone, R., Mohammed, A.K., and Hamon, C. (2003). Tandem

mass tags: a novel quantification strategy for comparative analysis of complex

protein mixtures by MS/MS. Anal. Chem. 75, 1895–1904.

Tsai, C.F., Hsu, C.C., Hung, J.N., Wang, Y.T., Choong, W.K., Zeng, M.Y., Lin,

P.Y., Hong, R.W., Sung, T.Y., and Chen, Y.J. (2014). Sequential phosphopro-

12 Cell Reports Methods 2, 100138, January 24, 2022

Xue, L., Geahlen, R.L., and Tao, W.A. (2013). Identification of direct tyrosine kinase substrates based on protein kinase assay-linked phosphoproteomics.

Mol. Cell. Proteomics 12, 2969–2980.

Yi, L., Tsai, C.F., Dirice, E., Swensen, A.C., Chen, J., Shi, T., Gritsenko, M.A.,

Chu, R.K., Piehowski, P.D., Smith, R.D., Rodland, K.D., Atkinson, M.A., Mathews, C.E., Kulkarni, R.N., Liu, T., and Qian, W.J. (2019). Boosting to amplify

signal with isobaric labeling (BASIL) strategy for comprehensive quantitative

phosphoproteomic characterization of small populations of cells. Anal.

Chem. 75, 5794–5801.

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Article

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE

SOURCE

IDENTIFIER

Chemicals, peptides, and recombinant proteins

Triethylammonium bicarbonate

Sigma

Catalog: T7408

Phosphatase Inhibitor Cocktail 2

Sigma

Catalog: P5726

Phosphatase Inhibitor Cocktail 3

Sigma

Catalog: P0044

BCA Protein Assay Kit

Thermo Scientific Pierce

Catalog: 23225

Sodium deoxycholate

FUJIFILM Wako

Catalog: 190-08313

Sodium lauroyl sarcosinate

FUJIFILM Wako

Catalog: 198-14745

Iron-(III) chloride

FUJIFILM Wako

Catalog: 091-00872

Dithiothreitol

Thermo Scientific

Catalog: 20291

Iodoacetamide

Thermo Scientific

Catalog: A3221

Lysyl endopeptidase

FUJIFILM Wako

Catalog: 129-02541

Sequencing-grade modified trypsin

Promega

Catalog: V517

Ni-NTA silica resins

QIAGEN

Catalog: 31314

Empore SDB-XC membrane disks

CDS

Catalog: 13-110-020

Titanium dioxide (10 mm)

GL Sciences

Catalog: 5020-75010

CK2a2/b (CSNK2A2/B)

Carna Biosciences

Catalog:05-185

PKACa(PRKACA)

Carna Biosciences

Catalog:01-127

ERK2 (MAPK1)

Carna Biosciences

Catalog:04-143

EGFR (ERBB1)

Carna Biosciences

Catalog:08-115

SRC

Carna Biosciences

Catalog:08-173

JNK1(MAPK8)

Carna Biosciences

Catalog:04-163

CDK1 (CDC2/CycB1)

Carna Biosciences

Catalog:04-102

p38a(MAPK14)

Carna Biosciences

Catalog:04-152

TMTsixplexTM

Thermo Scientific

Catalog:90061

https://zenodo.org/

https://doi.org/10.5281/zenodo.5750874

ATCC

Catalog: CCL-2.2

MaxQuant

PMID: 27809316

https://www.maxquant.org/

Perseus

PMID: 27348712

https://maxquant.net/perseus/

Deposited data

Zenodo

Experimental models: Cell lines

HeLa S3

Software and algorithms

RESOURCE AVAILABILITY

Lead contact

Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, Yasushi

Ishihama (yishiham@pharm.kyoto-u.ac.jp).

Materials availability

This study did not generate new unique reagents.

Data and code availability

Data described in this paper have been deposited at https://zenodo.org and are publicly available as of the date of publication.

DOIs are listed in the key resources table.

This paper does not report original code.

Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request

Cell Reports Methods 2, 100138, January 24, 2022 e1

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Article

EXPERIMENTAL MODEL AND SUBJECT DETAILS

HeLa S3 cells were cultured in DMEM containing 10% fetal bovine serum and 100 mg/mL kanamycin. For isobaric acidophilic motifcentric phosphoproteomes, cells were not stimulated (mock) or were stimulated with 10 mM CK2 inhibitor (CX-4945) for 30 min. For

isobaric basophilic motif-centric phosphoproteomes, cells were not stimulated (mock) or were stimulated with 10 mM PKA activator

(forskolin) for 30 min. For isobaric tyrosine and multiple motif-centric phosphoproteome, cells were treated with 10 mM EGF, 10 mM

EGF/10 mM afatinib, and 500 mM PV (pH 10 with 0.14% H2O2), respectively, for 30 min before harvesting. Two biological replicates

were performed.

METHOD DETAILS

Tryptic peptides from HeLa cell lysate

Cells were washed three times with ice-cold phosphate-buffered saline (phosphate-buffered saline, 0.01 M sodium phosphate,

0.14 M NaCl, pH 7.4) and then lysed in lysis buffer containing 12 mM sodium deoxycholate, 12 mM sodium lauroyl sarcosinate in

100 mM triethylammonium bicarbonate. Protein concentration was determined by means of BCA protein assay. The lysates were

digested based on the reported phase-transfer surfactants protocol (Masuda et al., 2008). The digested peptides were desalted

on SDB-XC StageTips (Rappsilber et al., 2007).

In vitro kinase reactions

For acidophilic, Pro-directed and tyrosine kinase reactions, the tryptic peptides were dissolved in 40 mM Tris-HCl (pH 7.5) and incubated with each kinase (0.2 mg CK2, ERK2, JNK1, p38a, CDK1 or SRC) at 37 C overnight for in vitro kinase reactions in the presence

of 1 mM ATP and 20 mM MgCl2. For the EGFR kinase reactions, tryptic peptides were firstly passed through SCX StageTips (Rappsilber et al., 2007) to remove afatinib. Eluted peptides were further desalted on SDB-XC StageTips. Then, the desalted peptides

were dissolved in 40 mM Tris-HCl (pH 7.5) and incubated with EGFR (0.2 mg) for in vitro kinase reactions in the presence of 1 mM

ATP and 4 mM MnCl2 at 37 C overnight. For basophilic kinases such as PKA, the lysates were loaded onto a 10-kDa ultrafiltration

device (Amicon Ultra, Millipore). The device was centrifuged at 14,000 g to remove the detergents. Subsequently, the original lysis

buffer was replaced with 40 mM Tris-HCl (pH 7.5) followed by centrifugation. Then, the proteins were incubated with 0.2 mg PKA for

in vitro kinase reactions in the presence of 1 mM ATP and 20 mM MgCl2 at 37 C overnight. After the kinase reactions, the proteins

were reduced with 10 mM DTT for 30 min at 37 C and alkylated with 50 mM iodoacetamide in the dark for 30 min at 37 C. The resulting samples were digested by Lys-C (1:100, w/w) at 37 C for 3 h followed by trypsin (1:50, w/w) overnight at 37 C. All the peptides

were desalted on SDB-XC StageTips.

TMT labeling for digested peptides

The desalted peptides were dissolved in 200 mM HEPES (pH 8.5). Then, the resuspended digested peptides were mixed with TMT

reagent dissolved in 100% ACN for 1 h. The labeling reaction was stopped by adding 5% hydroxylamine for 15 min, followed by acidification with TFA. All the peptides labeled with each multiplexed TMT reagent were mixed into the same tube and the mixture was

diluted to decrease the concentration of ACN to less than 5%. The TMT-labeled peptides were desalted on SDB-XC StageTips. The

information on the peptide amount in each TMT channel for all experiments is shown in Table S3. Note that the peptide amount for

each TMT channel was quantified by means of nanoLC-UV at 210 nm using a Thermo Ultimate 3000 RSLCnano system (Germering),

an MU701 UV detector (GL Sciences), and a C18 analytical column (150 mm length 3 100 mm ID) packed with Reprosil-Pur 120 C18AQ material (3 mm, Dr. Maisch).

IMAC

The procedure for phosphopeptides purification with an Fe3+-IMAC tip was as described previously (Tsai et al., 2014, 2015) with minor modifications. In brief, a buffer consisting of 50 mM EDTA in 1 M NaCl was used for removing Ni2+ ions. Then, the metal-free NTA

was activated by loading 100 mM FeCl3 into the IMAC tip. The Fe3+-IMAC tip was equilibrated with 0.5% (v/v) acetic acid at pH 3.0

before sample loading. Tryptic peptides from HeLa lysates were reconstituted in 0.5% (v/v) acetic acid and loaded onto the IMAC tip.

After successive washing steps with 1% (v/v) TFA in 80% ACN and 0.5% (v/v) acetic acid, the IMAC tip was coupled to an activated

SDB-XC StageTip and the bound phosphopeptides were eluted onto the SDB-XC StageTip with 200 mM NH4H2PO4 buffer. Then, the

eluted phosphopeptides were desalted with SDB-XC StageTip.

LC-MS/MS analysis

NanoLC-MS/MS analyses were performed on an Orbitrap Fusion Lumos Tribrid mass spectrometer (Thermo Scientific), which was

connected to the Thermo Ultimate 3000 RSLCnano system and an HTC-PAL autosampler (CTC Analytics). Peptide mixtures were

loaded onto and separated on self-pulled needle columns (150 mm length 3 100 mm inner diameter) packed with Reprosil-Pur

120 C18-AQ material (3 mm) or a 2-m-long C18 monolithic silica capillary column (Iwasaki et al., 2010). The mobile phases consisted

of (A) 0.5% acetic acid and (B) 0.5% acetic acid and 80% acetonitrile. Peptides were separated through a gradient from 17.5 % to

45% buffer B at a flow rate of 500 nL/min. Full-scan spectra were acquired at a target value of 43105 with a resolution of 60,000. Data

e2 Cell Reports Methods 2, 100138, January 24, 2022

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

Article

OPEN ACCESS

were acquired in a data-dependent acquisition mode using the top-speed method (3 s). The peptides were isolated using a quadrupole system (the isolation window was 0.7). The MS2 analysis was performed in the ion trap using collision-induced dissociation

fragmentation with a collision energy of 35 at a target value of 1 3 104 with 100 ms maximum injection time. The MS3 analysis was

performed for each MS2 scan acquired by using multiple MS2 fragment ions isolated by an ion trap as precursors for the MS3 analysis

with a multinotch isolation waveform (McAlister et al., 2014). HCD fragmentation was used for MS3 scan with an NCE of 65%, and the

fragment ions were detected by the Orbitrap (resolution 15,000). The AGC target was 53104 with a maximum ion injection time of

22 ms. The raw data sets have been deposited at the ProteomeXchange Consortium (http://proteomecentral.proteomexchange.

org) via the jPOST partner repository (https://jpostdb.org) (Moriya et al., 2019) with the dataset identifier JPST001027 (PXD026996).

Data analyses

Database search. The raw MS/MS data were processed with MaxQuant (Cox and Mann, 2008; Tyanova et al., 2016a). Peptide

search with full tryptic digestion and a maximum of two missed cleavages was performed against the SwissProt human database

(20,102 entries). The mass tolerance for precursor and MS3 ions was 4.5 ppm, whereas the tolerance for MS2 ions was 0.5 Th.

Acetylation (protein N-terminal), oxidation (M) and phospho (STY) were set as variable modifications and carbamidomethyl (C) was

set as a fixed modification. The quantitation function of reporter ion MS3 (6-plexed TMT) was turned on. The false discovery rate was

set to 1% at the level of PSMs and proteins. A score cut-off of 40 was used for identified modified peptides.

QUANTIFICATION AND STATISTICAL ANALYSIS

The abundances of TMT were log2-transformed and further analyzed by Perseus (Tyanova et al., 2016b) for statistical evaluation such

as principal component analyses and t tests. The PSP logo generator (Hornbeck et al., 2015) was used for sequence motif analysis.

DAVID (Huang da et al., 2009) was used for gene ontology and pathway enrichment analysis. STRING v11 (Szklarczyk et al., 2019)

was used for protein-protein interaction analysis. SigmaPlot (Systat Software), was used for preparing box plots.

Cell Reports Methods 2, 100138, January 24, 2022 e3

...

参考文献をもっと見る